This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

institutetext: InQubator for Quantum Simulation (IQuS), Department of Physics,
University of Washington, Seattle, WA 98195.

Entanglement minimization in hadronic scattering with pions

Silas R. Beane,    Roland C. Farrell, *    and Mira Varma***Address as of 1 June 2021: Department of Physics, Yale University.
(\mydate)
Abstract

Recent work Beane:2018oxh conjectured that entanglement is minimized in low-energy hadronic scattering processes. It was shown that the minimization of the entanglement power (EP) of the low-energy baryon-baryon SS-matrix implies novel spin-flavor symmetries that are distinct from large-Nc{N_{c}} QCD predictions and are confirmed by high-precision lattice QCD simulations. Here the conjecture of minimal entanglement is investigated for scattering processes involving pions and nucleons. The EP of the SS-matrix is constructed for the ππ\pi\pi and πN\pi N systems, and the consequences of minimization of entanglement are discussed and compared with large-Nc{N_{c}} QCD expectations.

dedication: NT@UW-21-05, IQuS@UW-21-008

1 Introduction

It is of current interest to uncover implications of quantum entanglement for the low-energy interactions of hadrons and nuclei111For a recent review, see Ref. Klco:2021lap .. As these interactions are profitably described by effective quantum field theory (EFT), which is an expansion of the relevant effective action in local operators, entanglement may have subtle implications for EFT which are difficult to identify due to its intrinsic non-locality. Ideally entanglement properties reveal themselves as regularities in hadronic data and, possibly, as accidental approximate symmetries. In addition to the non-local nature of entanglement, a difficulty lies with parsing the distinction, if any, between entanglement effects and generic quantum correlations which account for the deviation of QCD path integral configurations from a classical path. For instance, if one assumes that QCD with Nc=3{N_{c}}=3 is near the large-Nc{N_{c}} limit tHooft:1973alw ; Witten:1979kh ; Witten:1979pi ; Kaiser:2000gs , then one might expect that it would be difficult to distinguish between large-Nc{N_{c}} expectations and some fundamental underlying principle that minimizes entanglement independent of the value of Nc{N_{c}}. To make this more concrete, consider two local or non-local QCD operators 𝒪1{\cal O}_{1} and 𝒪2{\cal O}_{2}. If the vacuum expectation value of the product of these operators obeys the factorization rule tHooft:1973alw ; Witten:1979kh ; Witten:1979pi

𝒪1𝒪2=𝒪1𝒪2+O(ϵ)\langle{\cal O}_{1}{\cal O}_{2}\rangle\ =\ \langle{\cal O}_{1}\rangle\langle{\cal O}_{2}\rangle\ +\ O(\epsilon) (1)

where ϵ\epsilon is a small number, then the variance of any operator vanishes in the limit ϵ0\epsilon\rightarrow 0. A theory whose operators obey this factorization behaves like a classical theory222Ordinarily one identifies the classical theory with the trivial 0\hbar\rightarrow 0 limit. However, Ref. Yaffe:1981vf has established a more general criterion for the classical limit. and therefore has a small parameter ϵ\epsilon which measures quantum effects. Large-Nc{N_{c}} QCD is such a theory, and indeed, at least for a class of QCD operators, one can identify ϵ=1/Nc\epsilon=1/{N_{c}}. The factorization property, Eq. (1), is then easily deduced from Feynman diagrams involving quarks and gluons and amounts to the dominance of disconnected contributions in the path integral.

On the other hand, one might imagine that the factorization of Eq. (1) arises as a property of the path integral, rather than as a property of the local action (as in varying Nc{N_{c}} and taking it large in QCD). It is not a priori unlikely that, at least for a class of QCD operators, the path integral minimizes quantum fluctuations via a mechanism that is not currently understood. For instance, starting with QCD defined at short distances, the procedure of sequentially integrating out short distance modes to obtain low-energy hadronic scattering amplitudes may remove highly-entangled states that arise from non-perturbative QCD dynamics, leaving a low-energy EFT that is near a classical trajectory. It is intuitively sensible that the QCD confinement length acts as a natural cutoff of entanglement in the low-energy EFT. This notion can be raised to the conjecture that QCD will minimize the entanglement in low-energy hadronic interactions. Testing this conjecture relies on finding hadronic systems where its consequences deviate from those implied by large-Nc{N_{c}}. And the success of the large-Nc{N_{c}} approximation in describing the world renders this task challenging. Evidence in favor of this conjecture was found in Ref. Beane:2018oxh in a study of baryon-baryon scattering systems (See also Refs. Beane:2020wjl and Low:2021ufv ). This work relied both on theoretical arguments and high-precision lattice QCD simulations of baryon-baryon scattering systems with strangeness. In this paper, the conjecture of minimal entanglement will be investigated in both ππ\pi\pi and πN\pi N scattering.

Finding measures of the entanglement due to interaction is both non-trivial and non-unique. The most fundamental object in the scattering process is the unitary SS-matrix. In a scattering process in which the two in-state particles form a product state, the SS-matrix will entangle the in-state particles in a manner that is dependent on the energy of the scattering event. A useful measure of this entanglement is the entanglement power (EP) of the SS-matrix PhysRevA.63.040304 ; mahdavi2011cross ; Beane:2018oxh . In the case of nucleon-nucleon (NNNN) scattering, the EP was found for all momenta below inelastic threshold Beane:2018oxh . However, the most interesting phenomenological result is at threshold, where the vanishing EP implies the vanishing of the leading-order spin entangling operator, which in turn implies Wigner SU(4)SU(4) symmetry Wigner:1936dx ; Wigner:1937zz ; Wigner:1939zz . As this symmetry is a consequence of large-NcN_{c} QCD Kaplan:1995yg ; Kaplan:1996rk ; CalleCordon:2008cz , the minimization of entanglement and the large-NcN_{c} approximation are found to be indistinguishable in the two-flavor case. By contrast, in the three-flavor case, minimization of the entanglement power in baryon-baryon scattering implies an enhanced SU(16)SU(16) symmetry which is not necessarily implied by large-NcN_{c} and is realized in lattice QCD simulations Beane:2018oxh ; Low:2021ufv . Given that baryon-baryon scattering exhibits entanglement minimization, it is of interest to determine whether other low-energy QCD scattering systems exhibit this property. In investigating the EP of scattering systems involving pions, once again a crucial difficulty is distinguishing consequences of entanglement minimization and the large-Nc{N_{c}} limit. In the ππ\pi\pi system the implications of entanglement minimization are found to be indistinguishable from implications of large-NcN_{c}. In the πN\pi N system the implications of entanglement minimization are distinct, however the absence of an enhanced symmetry limits the predictive power to simple scaling laws with no smoking-gun predictions.

This paper is organized as follows. In Section 2, the EP of the ππ\pi\pi SS-matrix is considered in detail. After introducting the standard definition and conventions of the ππ\pi\pi SS-matrix, the SS-matrix is formulated in a basis convenient for calculation of the EP. Explicit expressions are derived for the EP of the first few partial waves in terms of phase shifts and leading-order expressions in chiral perturbation theory are provided. Using the highly-accurate Roy-equation solutions for the low-energy phase shifts, the experimental EP for the first few partial waves are given up to inelastic threshold. The consequences of minimizing the EP are considered and compared to large-Nc{N_{c}} expectations. In Section 3, the same procedure is repeated for the πN\pi N SS-matrix. Finally, Section 4 is a discussion of the possible conclusions that can be drawn from the conjecture of minimal entanglement.

2 The ππ\pi\pi System

There are, of course, several important differences between baryon-baryon and pion-pion scattering. Firstly, with pions there is no notion of spin entanglement. However, isospin (or generally flavor) entanglement is present and can be quantified using the EP and it is not clear that there is any meaningful distinction between these two kinds of entanglement. Indeed, it is straightforward to see that the “spin” entanglement of Ref. Beane:2018oxh can be reformulated as “isospin” entanglement with identical consequences333At the level of the EFT, this is simply realized via Fierz identities.. This is no surprise as entanglement is fundamentally a property of a non-product state vector whose existence relies either on an internal or a spacetime symmetry. Secondly, the crucial distinction between baryon-baryon scattering at very low-energies and the scattering of pions is that pion scattering at low-energies is strongly constrained by spontaneous chiral symmetry breaking in QCD. In particular, chiral symmetry implies that low-energy pion scattering on an arbitrary hadronic target is weak. The weak nature of the interaction is due to the smallness of the light-quark masses relative to a characteristic QCD scale. This translates to a chiral suppression of the EP at low-energies. Chiral symmetry breaking at large-Nc{N_{c}} does involve enhanced symmetry Kaiser:2000gs ; for NN flavors, the QCD chiral symmetries and their pattern of breaking are enhanced to U(N)U(N)U(N)U(N)\otimes U(N)\to U(N), as signaled by the presence of a new Goldstone boson, η\eta^{\prime}, whose squared mass scales as 1/Nc1/{N_{c}}. Intuitively, the anomaly, as an intrinsically quantum phenomenon, is a strongly entangling effect which would generally vanish as quantum fluctuations are suppressed. However, this is not assumed as the focus of this paper is two-body scattering which does not access the anomaly.

2.1 SS-matrix definition

The SS-matrix is defined as

S=1+iTS=1+iT (2)

where unity, corresponding to no interaction, has been separated out. This then defines the TT-matrix. The SS-matrix element for the scattering process πiπjπkπl\pi^{i}\pi^{j}\to\pi^{k}\pi^{l} is then given by

πk(p3)πl(p4)|S|πi(p1)πj(p2)=πk(p3)πl(p4)|πi(p1)πj(p2)+πk(p3)πl(p4)|iT|πi(p1)πj(p2)\langle\pi^{k}(p_{3})\pi^{l}(p_{4})|S|\pi^{i}(p_{1})\pi^{j}(p_{2})\rangle=\langle\pi^{k}(p_{3})\pi^{l}(p_{4})|\pi^{i}(p_{1})\pi^{j}(p_{2})\rangle\\ +\langle\pi^{k}(p_{3})\pi^{l}(p_{4})|iT|\pi^{i}(p_{1})\pi^{j}(p_{2})\rangle (3)

where ii, jj, kk, and ll are the isospin indices of the pion states. The projection operators onto states of definite isospin are444For a detailed construction, see Ref. lanz2018determination .

P0kl,ij\displaystyle P_{0}^{kl,ij} =\displaystyle= 13δklδij,\displaystyle\frac{1}{3}\delta^{kl}\delta^{ij}\ , (4)
P1kl,ij\displaystyle P_{1}^{kl,ij} =\displaystyle= 12(δkiδljδliδkj),\displaystyle\frac{1}{2}\left(\delta^{ki}\delta^{lj}-\delta^{li}\delta^{kj}\right)\ , (5)
P2kl,ij\displaystyle P_{2}^{kl,ij} =\displaystyle= 12(δkiδlj+δliδkj)13δklδij,\displaystyle\frac{1}{2}\left(\delta^{ki}\delta^{lj}+\delta^{li}\delta^{kj}\right)-\frac{1}{3}\delta^{kl}\delta^{ij}\ , (6)

where the subscript indicates the total isospin, II, of the ππ\pi\pi system. Straightforward field-theoretic manipulations then give

πk(p3)πl(p4)|S|πi(p1)πj(p2)=(2π)4δ4(p1+p2p3p4)16πσ(s)=0(2+1)P(cosθ)𝒮kl,ij,\langle\pi^{k}(p_{3})\pi^{l}(p_{4})|S|\pi^{i}(p_{1})\pi^{j}(p_{2})\rangle\\ =(2\pi)^{4}\delta^{4}(p_{1}+p_{2}-p_{3}-p_{4})\,\frac{16\pi}{\sigma(s)}\,\sum_{\ell=0}^{\infty}(2\ell+1)P_{\ell}(\cos\theta)\,{\bf\cal S}_{\ell}^{kl,ij}\ , (7)

where the PP_{\ell} are the Legendre polynomials, and

σ(s)14Mπ2/s,\sigma(s)\equiv\sqrt{1-4M_{\pi}^{2}/s}\ , (8)

with s=4(q2+Mπ2)s=4(q^{2}+M_{\pi}^{2}) and qq is the center-of-mass three-momentum of the pions. The focus here will be on the SS-matrices of definite partial wave:

𝒮kl,ije2iδ0P0kl,ij+e2iδ1P1kl,ij+e2iδ2P2kl,ij,{\bf\cal S}_{\ell}^{kl,ij}\equiv e^{2i\delta_{\ell}^{0}}P_{0}^{kl,ij}+e^{2i\delta_{\ell}^{1}}P_{1}^{kl,ij}+e^{2i\delta_{\ell}^{2}}P_{2}^{kl,ij}\ , (9)

which satisfy the unitarity constraint

𝒮kl,ij𝒮ij,mn=δkmδln.{\bf\cal S}_{\ell}^{kl,ij}{\bf\cal S}_{\ell}^{*ij,mn}\ =\ \delta^{km}\delta^{ln}\ . (10)

Since the pions obey Bose statistics, the angular momentum, \ell, is even for the states with I=0I=0 or 22 and odd for states with I=1I=1.

As the initial state in the scattering process is a product state of two pions, each in the 𝟑{\bf 3}-dimensional (I=1I=1) irrep of SU(2)SU(2) isospin, it is convenient to work in the direct-product matrix basis. The pion isospin matrices are the three-by-three matrices t^α\hat{t}_{\alpha} which satisfy

[t^α,t^β]=iϵαβγt^γ.\displaystyle{[\,\hat{t}_{\alpha}\,,\,\hat{t}_{\beta}\,]}\,=\,i\,\epsilon_{\alpha\beta\gamma}\,\hat{t}_{\gamma}\ . (11)

In the product Hilbert space 12{\cal H}_{1}\otimes{\cal H}_{2}, the total isospin of the two-pion system is 𝒕^13+3𝒕^2\hat{\bm{t}}_{1}\otimes{\cal I}_{3}+{\cal I}_{3}\otimes\hat{\bm{t}}_{2}, where 3{\cal I}_{3} is the three-by-three unit matrix, which implies

𝒕^1𝒕^2\displaystyle\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2} =\displaystyle= 12[I(I+1) 4]𝟏^=𝟏^{2,I=01,I=1  1,I=2\displaystyle\textstyle{1\over 2}\Big{[}I\left(I+1\right)\;-\;4\Big{]}\hat{\bf 1}\ =\ \hat{\bf 1}\begin{cases}-2,\qquad I=0\\ -1,\qquad I=1\\ \ \;1,\qquad I=2\end{cases} (12)

where 𝟏^=^3^3\hat{\bf 1}=\hat{\cal I}_{3}\otimes\hat{\cal I}_{3} and 𝒕^1𝒕^2=α=13t^1αt^2α\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}=\sum\limits_{\alpha=1}^{3}\ \hat{t}_{1}^{\alpha}\otimes\hat{t}_{2}^{\alpha}. The 9×99\times 9 dimensionality of the matrix is in correspondence with the dimensionality of the SU(2)SU(2) isospin product representation 𝟑𝟑=𝟏𝟑𝟓{\bf 3}\otimes{\bf 3}={\bf 1}\oplus{\bf 3}\oplus{\bf 5}. There are now three invariants and three observables; one easily finds the SS-matrix in the direct-product matrix basis

𝐒^\displaystyle\hat{\bf S}_{\ell} =\displaystyle= e2iδ0𝐏^0+e2iδ1𝐏^1+e2iδ2𝐏^2,\displaystyle e^{2i\delta_{\ell}^{0}}\hat{\bf P}_{0}+e^{2i\delta_{\ell}^{1}}\hat{\bf P}_{1}+e^{2i\delta_{\ell}^{2}}\hat{\bf P}_{2}\ , (13)

where the three 9×99\times 9 projection matrices are

𝐏^0\displaystyle\hat{\bf P}_{0} =\displaystyle= 13(𝟏^(𝒕^1𝒕^2)2),\displaystyle-\frac{1}{3}\left(\hat{\bf 1}-\left(\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}\right)^{2}\right)\ , (14)
𝐏^1\displaystyle\hat{\bf P}_{1} =\displaystyle= 𝟏^12((𝒕^1𝒕^2)+(𝒕^1𝒕^2)2),\displaystyle\hat{\bf 1}-\frac{1}{2}\left(\left(\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}\right)+\left(\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}\right)^{2}\right)\ , (15)
𝐏^2\displaystyle\hat{\bf P}_{2} =\displaystyle= 13(𝟏^+32(𝒕^1𝒕^2)+12(𝒕^1𝒕^2)2).\displaystyle\frac{1}{3}\left(\hat{\bf 1}+\frac{3}{2}\left(\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}\right)+\frac{1}{2}\left(\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}\right)^{2}\right)\ . (16)

It is readily checked that the SS-matrix is unitary, and using the representation (tγ)αβ=iϵαβγ({t}_{\gamma})_{\alpha\beta}=-i\epsilon_{\alpha\beta\gamma}, it is straightforward to establish equivalence with the component form, Eq. (9). The trace is given by ei2δ0+3ei2δ1+5ei2δ2e^{i2\delta_{\ell}^{0}}+3e^{i2\delta_{\ell}^{1}}+5e^{i2\delta_{\ell}^{2}} which correctly counts the isospin multiplicity, and is in correspondence with the nine eigenvalues of 𝐒^\hat{\bf S}.

2.2 Entanglement power

Consider the =1\ell=1 SS-matrix. As this system can scatter only in the I=1I=1 channel, it provides a useful example of how the SS-matrix entangles the initial two-pion state. From Eq. (13) one finds

𝐒^1\displaystyle\hat{\bf S}_{1} =\displaystyle= 12(1+ei2δ11)𝟏^+12(1ei2δ11)𝒫12\displaystyle{1\over 2}\left(1+e^{i2\delta_{1}^{1}}\right)\hat{\bf 1}\ +\ {1\over 2}\left(1-e^{i2\delta_{1}^{1}}\right){\cal P}_{12} (17)

where the SWAP operator is given by

𝒫12=(𝒕^1𝒕^2)2+𝒕^1𝒕^2𝟏^.{\cal P}_{12}=\left(\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}\right)^{2}+\hat{\bm{t}}_{1}\cdot\hat{\bm{t}}_{2}-\hat{\bf 1}\ . (18)

As the SWAP operator interchanges the pions in the initial two-pion product state, leaving another two-pion product state, it does not entangle. Therefore, the SS-matrix has the two obvious non-entangling solutions δ11=0\delta_{1}^{1}=0 (no interaction) and δ11=π/2\delta_{1}^{1}=\pi/2 (at resonance). One measure of SS-matrix entanglement would then be the (absolute value squared of the) product of the coefficients of the non-entangling solutions:

|(1+ei2δ11)(1ei2δ11)|2sin2(2δ11).\displaystyle\Big{\lvert}\;\left(1+e^{i2\delta_{1}^{1}}\right)\left(1-e^{i2\delta_{1}^{1}}\right)\Big{\lvert}^{2}\;\sim\sin^{2}\left(2\delta_{1}^{1}\right)\ . (19)

A state-independent measure of the entanglement generated by the action of the SS-matrix on the initial product state of two free particles is the EP PhysRevA.63.040304 ; mahdavi2011cross ; Beane:2018oxh . In order to compute the EP an arbitrary initial product state should be expressed in a general way that allows averaging over a given probability distribution folded with the initial state. Recall that in the NNNN case, there are two spin states (a qubit) for each nucleon and therefore the most general initial nucleon state involves two complex parameters or four real parameters. Normalization gets rid of one parameter and there is an overall irrelevant phase which finally leaves two real parameters which parameterize the 𝐂𝐏1{\bf CP}^{1} manifold, also known as the 2-sphere 𝐒2{\bf S}^{2}, or the Bloch sphere. Now in the isospin-one case we have three isospin states (a qutrit) which involves three complex parameters. Again normalization and removal of the overall phase reduce this to four real parameters which parameterize the 𝐂𝐏2{\bf CP}^{2} manifold Brody_2001 ; Bengtsson:2001yd ; bengtsson_zyczkowski_2006 . Since the ππ\pi\pi initial state is the product of two isospin-one states, there will be eight parameters to integrate over to get the EP.

There are now two qutrits in the initial state, which live in the Hilbert spaces 1,2{\cal H}_{1,2}, each spanned by the states {|𝟏i,| 0i,| 1i}\{|\,{\bf-1}_{i}\,\rangle,|\,{\bf 0}_{i}\,\rangle,|\,{\bf 1}_{i}\,\rangle\} with i=1,2i=1,2. It is of interest to determine the EP of a given SS-matrix operator, which is a measure of the entanglement of the scattered state averaged over the 𝐂𝐏2{\bf CP}^{2} manifolds on which the qutrits live. Consider an arbitrary initial product state of the qutrits

|Ψ=U(α1,β1,μ1,ν1)|1U(α2,β2,μ2,ν2)|2\displaystyle|\,\Psi\,\rangle\ =\ U\left(\alpha_{1},\beta_{1},\mu_{1},\nu_{1}\right)|\,\rangle_{1}\otimes U\left(\alpha_{2},\beta_{2},\mu_{2},\nu_{2}\right)|\,\rangle_{2} (20)

with

U(αi,βi,μi,νi)|i=cosβisinαi|𝟏i+eiμisinβisinαi| 0i+eiνicosαi| 1i,\displaystyle U\left(\alpha_{i},\beta_{i},\mu_{i},\nu_{i}\right)|\,\rangle_{i}=\cos\beta_{i}\sin\alpha_{i}|\,{\bf-1}\,\rangle_{i}+e^{i\mu_{i}}\sin\beta_{i}\sin\alpha_{i}|\,{\bf 0}\,\rangle_{i}+e^{i\nu_{i}}\cos\alpha_{i}|\,{\bf 1}\,\rangle_{i}\,, (21)

where 0μi,νi<2π0\leq{\mu_{i},\nu_{i}}<2{\pi} and 0αi,βiπ/20\leq{\alpha_{i},\beta_{i}}\leq{\pi}/2. The geometry of 𝐂𝐏2{\bf CP}^{2} is described by the Fubini-Study (FS) line element Brody_2001 ; Bengtsson:2001yd ; bengtsson_zyczkowski_2006

dsFS2=dα2+sin2(α)dβ2+(sin2(α)sin2(β)sin4(α)sin4(β))dμ2+sin2(α)cos2(α)dν22sin2(α)cos2(α)sin2(β)dμdν.\displaystyle\begin{split}ds_{\scriptstyle FS}^{2}=\;&d\alpha^{2}+\sin^{2}(\alpha)d\beta^{2}+\left(\sin^{2}(\alpha)\sin^{2}(\beta)-\sin^{4}(\alpha)\sin^{4}(\beta)\right)d\mu^{2}+\\ &\sin^{2}(\alpha)\cos^{2}(\alpha)d\nu^{2}-2\sin^{2}(\alpha)\cos^{2}(\alpha)\sin^{2}(\beta)d\mu d\nu\ .\end{split} (22)

Of special interest here is the differential volume element which in these coordinates is

dVFS\displaystyle dV_{\scriptstyle FS} =\displaystyle= detgFSdαdβdμdν\displaystyle\sqrt{det\,g_{\scriptstyle FS}}\,d\alpha\,d\beta\,d\mu\,d\nu (23)
=\displaystyle= cosαcosβsin3αsinβdαdβdμdν\displaystyle\cos\alpha\cos\beta\sin^{3}\alpha\sin\beta\,d\alpha\,d\beta\,d\mu\,d\nu\

and the volume of the 𝐂𝐏2{\bf CP}^{2} manifold is found to be,

𝑑VFS\displaystyle\int dV_{\scriptstyle FS} =\displaystyle= π22.\displaystyle\frac{\pi^{2}}{2}\ . (24)

The final state of the scattering process is obtained by acting with the unitary SS-matrix of definite angular momentum on the arbitrary initial product state:

|Ψ¯=𝐒^|Ψ.\displaystyle|\,\bar{\Psi}\,\rangle\ =\ \hat{\bf S}_{\ell}|\,\Psi\,\rangle\ . (25)

The associated density matrix is

ρ1,2=|Ψ¯Ψ¯|,\displaystyle\rho_{1,2}\ =\ |\,\bar{\Psi}\,\rangle\langle\,\bar{\Psi}|\,, (26)

and tracing over the states in 2{\cal H}_{2} gives the reduced density matrix

ρ1=Tr2[ρ1,2].\displaystyle\rho_{1}\ =\ {\rm Tr}_{2}\big{[}\rho_{1,2}\big{]}. (27)

The linear entropy of the SS-matrix is then defined as555Note that this is related to the (exponential of the) second Rényi entropy.

E𝐒^= 1Tr1[(ρ1)2].\displaystyle E_{\hat{\bf S}_{\ell}}\ =\ 1\ -\ {\rm Tr}_{1}\big{[}\left(\rho_{1}\right)^{2}\big{]}. (28)

Finally, the linear entropy is integrated over the initial 𝐂𝐏2{\bf CP}^{2} manifolds to form the average, and the entanglement power is

(𝐒^)=(2π2)2(i=12𝑑VFSi)𝒫E𝐒^\displaystyle{\mathcal{E}}({\hat{\bf S}_{\ell}})\ =\ \left(\frac{2}{\pi^{2}}\right)^{2}\left(\prod_{i=1}^{2}\int dV^{i}_{\scriptstyle FS}\right){\cal P}E_{\hat{\bf S}_{\ell}} (29)

where 𝒫{\cal P} is a probability distribution which here will be taken to be unity. Evaluating this expression using Eq. (13) yields the s-wave ππ\pi\pi EP:

(𝐒^0)\displaystyle{\mathcal{E}}({\hat{\bf S}_{0}}) =\displaystyle= 1648(1566cos[4δ00]65cos[2(δ00δ02)]\displaystyle\frac{1}{648}\left(156-6\cos[4\delta_{0}^{0}]-65\cos[2(\delta_{0}^{0}-\delta_{0}^{2})]\right. (30)
10cos[4(δ00δ02)]60cos[4δ02]15cos[2(δ00+δ02)]),\displaystyle\qquad\qquad\qquad\left.-10\cos[4(\delta_{0}^{0}-\delta_{0}^{2})]-60\cos[4\delta_{0}^{2}]-15\cos[2(\delta_{0}^{0}+\delta_{0}^{2})]\right)\ ,

and the p-wave ππ\pi\pi EP:

(𝐒^1)=14sin2(2δ11).\displaystyle{\mathcal{E}}({\hat{\bf S}_{1}})=\frac{1}{4}\sin^{2}\left(2\delta_{1}^{1}\right)\ . (31)

Notice that this matches the intuitive construction which led to Eq. (19). The EPs have the non-entangling solutions:

δ00\displaystyle\delta_{0}^{0} =\displaystyle= δ02= 0,π2,\displaystyle\delta_{0}^{2}\ =\ 0,\frac{\pi}{2}\ , (32)
δ11\displaystyle\delta_{1}^{1} =\displaystyle= 0,π2.\displaystyle 0,\frac{\pi}{2}\ . (33)

Therefore, in the s-wave, entanglement minimization implies that both isospins are either non-interacting or at resonance, while in the p-wave, entanglement minimization implies that the I=1I=1 channel is either non-interacting or at resonance. As no I=2I=2 resonances are observed in nature (and their coupling to pions is suppressed in large-Nc{N_{c}} QCD Weinberg:2013cfa ), the s-wave EP has a single minimum corresponding to no interaction. By contrast, the I=1I=1 channel will exhibit minima of both types. It is worth considering the EP of a simple resonance model. Consider the unitary SS-matrix:

𝐒^1=sm12im1Γ1sm12+im1Γ1,\displaystyle{\hat{\bf S}_{1}}\ =\ \frac{s-m_{1}^{2}-im_{1}\Gamma_{1}}{s-m_{1}^{2}+im_{1}\Gamma_{1}}\ , (34)

where m1m_{1} (Γ1\Gamma_{1}) are the mass (width) of the resonance. The EP is

(𝐒^1)=(m1Γ1(sm12)(m1Γ1)2+(sm12)2)2,\displaystyle{\mathcal{E}}({\hat{\bf S}_{1}})=\left(\frac{m_{1}\Gamma_{1}\left(s-m_{1}^{2}\right)}{\left(m_{1}\Gamma_{1}\right)^{2}+\left(s-m_{1}^{2}\right)^{2}}\right)^{2}\ , (35)

which vanishes on resonance at s=m12s=m_{1}^{2} and has maxima at s=m1(m1±Γ1)s=m_{1}(m_{1}\pm\Gamma_{1}). It is clear that the minimum corresponds to 𝐒^𝒫12\hat{\bf S}\propto{\cal P}_{12}. As the ρ\rho-resonance dominates the I=1I=1 channel at energies below 1GeV1~{}{\rm GeV}, the EP in nature will be approximately of this form.

The ππ\pi\pi phase shifts are the most accurately known of all hadronic SS-matrices as the Roy equation constraints Roy:1971tc come very close to a complete determination of the phase shifts Ananthanarayan:2000ht ; Colangelo:2001df . In Fig. (1) the EPs for the first few partial waves are plotted using the Roy equation determinations of the SS-matrix.

Refer to caption
Figure 1: Entanglement power of the ππ\pi\pi SS-matrix for =0,1\ell=0,1 taken from Roy equation determinations (the bands represent an estimate of the uncertainties Ananthanarayan:2000ht ; Colangelo:2001df ) of the ππ\pi\pi phase shifts.

2.3 Chiral perturbation theory

Near threshold, the phase shift can be expressed in the effective range expansion as

δI(s)=12sin1{2σ(s)q2(aI+𝒪(q2))},\delta_{\ell}^{I}(s)=\textstyle{1\over 2}\sin^{-1}\{2\sigma(s)q^{2\ell}\left(a^{I}_{\ell}\;+\;\mathcal{O}(q^{2})\right)\}\ , (36)

where the scattering lengths, aIa^{I}_{\ell}, relevant to s-wave and p-wave scattering, are given at leading order in chiral perturbation theory by Weinberg:1978kz ; Gasser:1983yg

a00=7Mπ232πFπ2,a02=Mπ216πFπ2,a11=124πFπ2,a_{0}^{0}\ =\ \frac{7M_{\pi}^{2}}{32\pi F_{\pi}^{2}}\ \ ,\ \ a_{0}^{2}\ =\ -\frac{M_{\pi}^{2}}{16\pi F_{\pi}^{2}}\ \ ,\ \ a_{1}^{1}\ =\ \frac{1}{24\pi F_{\pi}^{2}}\ , (37)

where Fπ=93MeVF_{\pi}=93~{}{\rm MeV} is the pion decay constant. Near threshold the s-wave and p-wave EPs are given by

(𝐒^0)=19Mπ2[4(a00)25(a00a02)+10(a02)2]q2+𝒪(q4),\displaystyle{\mathcal{E}}({\hat{\bf S}_{0}})\ =\ \frac{1}{9M_{\pi}^{2}}\big{[}4(a^{0}_{0})^{2}-5(a^{0}_{0}a^{2}_{0})+10(a^{2}_{0})^{2}\big{]}\;q^{2}\ +\ \mathcal{O}(q^{4})\ ,
(𝐒^1)=1Mπ2(a11)2q6+𝒪(q8).\displaystyle{\mathcal{E}}({\hat{\bf S}_{1}})\ =\ \frac{1}{M_{\pi}^{2}}(a^{1}_{1})^{2}\;q^{6}\ +\ \mathcal{O}(q^{8})\ . (38)

As a00a_{0}^{0} (a02a_{0}^{2} ) is positive (negative) definite, the EP is trivially minimized with vanishing scattering lengths. This then implies a bookkeeping where Fπ=𝒪(ϵn)F_{\pi}=\mathcal{O}(\epsilon^{-n}) where nn is a positive number. Hence, in the limit of vanishing entanglement, the pions are non-interacting, and the dominant interaction is from tree diagrams; i.e. loops are suppressed by inverse powers of FπF_{\pi}. In the large-Nc{N_{c}} limit, one finds ϵ=1/Nc\epsilon=1/N_{c} and n=1/2n=1/2 tHooft:1973alw ; Witten:1979kh ; Witten:1979pi . Evidently the implications of vanishing entanglement for the ππ\pi\pi SS-matrix are indistinguishable from large-Nc{N_{c}} expectations666We also studied the effect of explicit chiral symmetry breaking on the entanglement power by varying the coefficients of operators with insertions of the quark mass matrix in the effective action. No evidence of a connection between chiral symmetry breaking and the entanglement power was found. This aligns with large-Nc{N_{c}} expectations as the meson masses are independent of Nc{N_{c}}. For an example of a relationship between entanglement and chiral symmetry breaking see Beane:2019loz ..

3 The πN\pi N System

As baryons are formed from Nc{N_{c}} quarks, the baryon masses and axial matrix elements grow with Nc{N_{c}}. The unitarity of the SS-matrix then places powerful constraints on baryon properties via large-Nc{N_{c}} consistency conditions Dashen:1993as ; Dashen:1993ac ; Dashen:1993jt ; Dashen:1994qi . At leading order in the large-Nc{N_{c}} expansion this yields predictions that are equivalent in the two (three) flavor case to SU(4)SU(4) (SU(6)SU(6)) spin-flavor symmetry which place the ground-state baryon spin states in the 𝟐𝟎{\bf 20} (𝟓𝟔{\bf 56}) dimensional irrep together with the delta (baryon decuplet). Therefore, the large-Nc{N_{c}} limit not only predicts an enhanced symmetry but also alters the definition of a baryon in a fundamental way. Moreover, any sensible effective theory of πN\pi N scattering in the large-Nc{N_{c}} limit must include the delta resonance as an explicit degree of freedom. In what follows, the consequences of entanglement minimization of the low-energy SS-matrix are considered for Nc=3{N_{c}}=3 QCD.

3.1 SS-matrix definition

The SS-matrix element for the scattering process, πa(q1)N(p1)πb(q2)N(p2)\pi^{a}(q_{1})N(p_{1})\to\pi^{b}(q_{2})N(p_{2}), is given by

πb(q2)N(p2)|S|πa(q1)N(p1)=πb(q2)N(p2)|πa(q1)N(p1)+πb(q2)N(p2)|iT|πa(q1)N(p1),\langle\pi^{b}(q_{2})N(p_{2})|S|\pi^{a}(q_{1})N(p_{1})\rangle=\langle\pi^{b}(q_{2})N(p_{2})|\pi^{a}(q_{1})N(p_{1})\rangle\\ +\langle\pi^{b}(q_{2})N(p_{2})|iT|\pi^{a}(q_{1})N(p_{1})\rangle, (39)

where aa and bb label the isospin of the pion. The TT matrix element in the center-of-mass system (cms) for the process may be parameterized as Fettes_1998

TπNba=(E+m2m){δba[g+(ω,t)+iσ(q2×q1)h+(ω,t)]+iϵabcτc[g(ω,t)+iσ(q2×q1)h(ω,t)]}\begin{split}T^{ba}_{\pi N}=\left({E+m\over 2m}\right)\bigg{\{}\delta^{ba}&\left[g^{+}(\omega,t)+i\vec{\sigma}\cdot(\vec{q_{2}}\times\vec{q_{1}})h^{+}(\omega,t)\right]\\ +i\epsilon^{abc}\tau^{c}&\left[g^{-}(\omega,t)+i\vec{\sigma}\cdot(\vec{q_{2}}\times\vec{q_{1}})h^{-}(\omega,t)\right]\bigg{\}}\end{split} (40)

where EE is the nucleon energy, ω\omega is the pion energy, mm is the nucleon mass and t=(q1q2)2t=(q_{1}-q_{2})^{2} is the square of the momentum transfer. The σ\sigma(τ\tau) matrices act on the spin(isospin) of the incoming nucleon. This decomposition reduces the scattering problem to calculating g±g^{\pm}, the isoscalar/isovector non-spin-flip amplitude and h±h^{\pm}, the isoscalar/isovector spin-flip amplitude. The amplitude can be further projected onto partial waves by integrating against PP_{\ell}, the relevant Legendre polynomial:

f±±(s)=E+m16πs1+1𝑑z[g±P(z)+q 2h±(P±1(z)zP(z))].f_{\ell\pm}^{\pm}(s)={E+m\over 16\pi\sqrt{s}}\int_{-1}^{+1}dz\left[g^{\pm}P_{\ell}(z)+{\vec{q}}^{\,2}h^{\pm}\big{(}P_{\ell\pm 1}(z)-zP_{\ell}(z)\big{)}\right]\ . (41)

Here z=cosθz=\cos{\theta} is the cosine of the scattering angle, ss is the cms energy squared and q 2=q12=q22\vec{q}^{\,2}=\vec{q_{1}}^{2}=\vec{q_{2}}^{2}. The subscript ±\pm on the partial wave amplitude indicates the total angular momentum J=±sJ=\ell\pm s. The amplitudes in the total isospin I=12I=\frac{1}{2} and I=32I=\frac{3}{2} can be recovered via the identification:

f±12=f±++2f±,f±32=f±+f±.f_{\ell\pm}^{\frac{1}{2}}=f_{\ell\pm}^{+}+2f_{\ell\pm}^{-}\ \ ,\ \ f_{\ell\pm}^{\frac{3}{2}}=f_{\ell\pm}^{+}-f_{\ell\pm}^{-}\ . (42)

Below inelastic threshold the scattering amplitude is related to a unitary SS-matrix by

S±I(s)=1+2i|q|f±I(s),S±I(s)S±I(s)=1S_{\ell\pm}^{I}(s)=1+2i\lvert\,\vec{q}\,\rvert f_{\ell\pm}^{I}(s)\ \ ,\ \ S_{\ell\pm}^{I}(s)S_{\ell\pm}^{I}(s)^{\dagger}=1 (43)

and the SS-matrix can be parameterized in terms of phase shifts,

S±I(s)=e2iδ±I(s).S_{\ell\pm}^{I}(s)=e^{2i\delta_{\ell\pm}^{I}(s)}\ . (44)

For a more detailed derivation of the πN\pi N SS-matrix see Fettes_1998 ; Yao_2016 ; Moj_i__1998 ; Scherer2012 . Scattering in a given partial wave and total angular momentum channel leads to a SS-matrix which acts on the product Hilbert space of the nucleon and pion isospin, πN\mathcal{H}_{\pi}\otimes\mathcal{H}_{N}. The SS-matrix can then be written in terms of total isospin projection operators

𝐒^±=e2iδ±3/2𝐏^3/2+e2iδ±1/2𝐏^1/2\hat{{\bf S}}_{\ell\pm}=e^{2i\delta_{\ell\pm}^{3/2}}\hat{{\bf P}}_{3/2}+e^{2i\delta_{\ell\pm}^{1/2}}\hat{{\bf P}}_{1/2} (45)

where the 6×66\times 6 projection matrices are

𝐏^3/2=23(𝟏^+𝐭^π𝐭^N),𝐏^1/2=13(𝟏^2(𝐭^π𝐭^N)).\begin{split}&\hat{{\bf P}}_{3/2}=\frac{2}{3}\left(\hat{{\bf 1}}+\hat{{\bf t}}_{\pi}\cdot\hat{{\bf t}}_{N}\right)\ ,\\ &\hat{{\bf P}}_{1/2}=\frac{1}{3}\left(\hat{{\bf 1}}-2(\hat{{\bf t}}_{\pi}\cdot\hat{{\bf t}}_{N})\right)\ .\end{split} (46)

The operators 𝐭^N\hat{{\bf t}}_{N} and 𝐭^π\hat{{\bf t}}_{\pi} are in the 22 and 33 dimensional representations of SU(2)SU(2) isospin respectively and 𝒕^π𝒕^N=α=13t^παt^Nα\hat{\bm{t}}_{\pi}\cdot\hat{\bm{t}}_{N}=\sum\limits_{\alpha=1}^{3}\ \hat{t}_{\pi}^{\alpha}\otimes\hat{t}_{N}^{\alpha}.

3.2 Entanglement power

The entanglement power of the πN\pi N SS-matrix can be computed in a similar manner as for the ππ\pi\pi EP. The incoming separable state now maps to a point on the product manifold, 𝐂𝐏2×𝐒2{\bf CP}^{2}\times{\bf S}^{2}. The construction of the reduced density matrix follows the same steps as in section 2.2 and the entanglement power is found to be,

(𝐒^±)=(2π214π)(𝑑VFS𝑑Ω)𝒫E𝐒^±=8243[17+10cos(2(δ±3/2δ±1/2))]sin2(δ±3/2δ±1/2)\begin{split}{\mathcal{E}}({\hat{\bf S}}_{\ell\pm})&=\left(\frac{2}{\pi^{2}}\frac{1}{4\pi}\right)\left(\int dV_{FS}d\Omega\right)\mathcal{P}E_{\hat{\bf S}_{\ell\pm}}\\ &=\frac{8}{243}\bigg{[}17+10\cos\left(2\big{(}\delta_{\ell\pm}^{{3/2}}-\delta_{\ell\pm}^{{1/2}}\big{)}\right)\bigg{]}\sin^{2}\left(\delta_{\ell\pm}^{{3/2}}-\delta_{\ell\pm}^{{1/2}}\right)\end{split} (47)

where 𝒫\mathcal{P} has been taken to be 11. Note that the two particles are now distinguishable and so scattering in each partial wave is no longer constrained by Bose/Fermi statistics. It follows that the SS-matrix is only non-entangling when it is proportional to the identity which occurs when,

δ±3/2=δ±1/2.\delta_{\ell\pm}^{{3/2}}=\delta_{\ell\pm}^{{1/2}}\ . (48)

Notice that the EP allows for interesting local minima when the difference in I=3/2I=3/2 and I=1/2I=1/2 phase shifts is π/2\pi/2. The πN\pi N phase shifts are determined very accurately by the Roy-Steiner equations up to a center-of-mass energy of 1.38 GeV Hoferichter_2016 and the entanglement power for the first couple partial waves is shown in Fig. (2).

Refer to caption
Figure 2: Entanglement power of the πN\pi N SS-matrix for =0,1\ell=0,1 taken from Roy-Steiner equation determinations (the bands represent an estimate of the uncertainties Hoferichter_2016 ) of the πN\pi N phase shifts.

There is a local minimum near the delta resonance position in the p-wave due to the rapid change of the I=3/2I=3/2 phase shift.

3.3 Chiral perturbation theory

Near threshold the phase shifts can be determined by the scattering lengths through the effective range expansion,

δ±I=cot1{1|q|2+1(1a±I+𝒪(q 2))}.\delta_{\ell\pm}^{I}=\cot^{-1}{\left\{{1\over\lvert\,\vec{q}\,\rvert^{2\ell+1}}\left({1\over a_{\ell\pm}^{I}}+\mathcal{O}(\vec{q}^{\,2})\right)\right\}}\ . (49)

This leads to the threshold form of the entanglement power,

(𝐒^±)=89(a±12a±32)2q 2+4l{\mathcal{E}}({\hat{\bf S}}_{\ell\pm})=\frac{8}{9}\left(a^{1\over 2}_{\ell\pm}-a^{3\over 2}_{\ell\pm}\right)^{2}\vec{q}^{\,2+4l} (50)

which can only vanish if a±12=a±32a^{1\over 2}_{\ell\pm}=a^{3\over 2}_{\ell\pm}. The scattering lengths at leading order in heavy-baryon chiral perturbation theory, including the delta, are given by Fettes_1998 ; Fettes_2001 ,

a0+12=2Mπm8π(m+Mπ)Fπ2,a0+32=Mπm8π(m+Mπ)Fπ2a112=m(9gA2Δ+9gA2Mπ8gπNΔ2Mπ)54πFπ2Mπ(Δ+Mπ)(m+Mπ),a132=m(9gA2Δ+9gA2Mπ8gπNΔ2Mπ)216πFπ2Mπ(Δ+Mπ)(m+Mπ)a1+12=m(3gA2Δ+3gA2Mπ+8gπNΔ2Mπ)72πFπ2Mπ(ΔMπ)(m+Mπ),a1+32=m(3gA2Δ22gπNΔ2MπΔ+3gA2Mπ2)36πFπ2Mπ(Mπ2Δ2)(m+Mπ)\begin{split}&a^{1\over 2}_{0+}=\frac{2M_{\pi}m}{8\pi(m+{M_{\pi}})F_{\pi}^{2}}\ \ ,\ \ a^{3\over 2}_{0+}=\frac{-M_{\pi}m}{8\pi(m+{M_{\pi}})F_{\pi}^{2}}\\ &a^{1\over 2}_{1-}=-\frac{m\left(9g_{A}^{2}\Delta+9g_{A}^{2}M_{\pi}-8g_{\pi N\Delta}^{2}M_{\pi}\right)}{54\pi F_{\pi}^{2}M_{\pi}(\Delta+M_{\pi})(m+M_{\pi})}\ \ ,\ \ a^{3\over 2}_{1-}=-\frac{m\left(9g_{A}^{2}\Delta+9g_{A}^{2}M_{\pi}-8g_{\pi N\Delta}^{2}M_{\pi}\right)}{216\pi F_{\pi}^{2}M_{\pi}(\Delta+M_{\pi})(m+M_{\pi})}\\ &a^{1\over 2}_{1+}=\frac{m\left(-3g_{A}^{2}\Delta+3g_{A}^{2}M_{\pi}+8g_{\pi N\Delta}^{2}M_{\pi}\right)}{72\pi F_{\pi}^{2}M_{\pi}(\Delta-M_{\pi})(m+M_{\pi})}\ \ ,\ \ a^{3\over 2}_{1+}=\frac{m\left(-3g_{A}^{2}\Delta^{2}-2g_{\pi N\Delta}^{2}M_{\pi}\Delta+3g_{A}^{2}M_{\pi}^{2}\right)}{36\pi F_{\pi}^{2}M_{\pi}\left(M_{\pi}^{2}-\Delta^{2}\right)(m+M_{\pi})}\end{split} (51)

where Δ=mΔmN\Delta=m_{\Delta}-m_{N} is the delta-nucleon mass splitting. The corresponding EPs near threshold are,

(𝐒^0+)=m2Mπ28π2Fπ4(m+Mπ)2q 2(𝐒^1)=m2(9gA2Δ+9gA2Mπ8gπNΔ2Mπ)25832π2fπ4Mπ2(Δ+Mπ)2(m+Mπ)2q 6(𝐒^1+)=m2(9gA2Δ2+4gπNΔ2ΔMπ+(9gA2+8gπNΔ2)Mπ2)25832π2fπ4Mπ2(Mπ2Δ2)2(m+Mπ)2q 6.\begin{split}&{\mathcal{E}}({\hat{\bf S}}_{0+})=\frac{m^{2}M_{\pi}^{2}}{8\pi^{2}F_{\pi}^{4}(m+M_{\pi})^{2}}\vec{q}^{\,2}\\ &{\mathcal{E}}({\hat{\bf S}}_{1-})=\frac{m^{2}\left(9g_{A}^{2}\Delta+9g_{A}^{2}M_{\pi}-8g_{\pi N\Delta}^{2}M_{\pi}\right)^{2}}{5832\pi^{2}f_{\pi}^{4}M_{\pi}^{2}(\Delta+M_{\pi})^{2}(m+M_{\pi})^{2}}\vec{q}^{\,6}\\ &{\mathcal{E}}({\hat{\bf S}}_{1+})=\frac{m^{2}\left(-9g_{A}^{2}\Delta^{2}+4g_{\pi N\Delta}^{2}\Delta M_{\pi}+\left(9g_{A}^{2}+8g_{\pi N\Delta}^{2}\right)M_{\pi}^{2}\right)^{2}}{5832\pi^{2}f_{\pi}^{4}M_{\pi}^{2}\left(M_{\pi}^{2}-\Delta^{2}\right)^{2}(m+M_{\pi})^{2}}\vec{q}^{\,6}\ .\end{split} (52)

Once again the only non-entangling solution consistent with chiral symmetry is no interaction, with the same scaling of FπF_{\pi} as found in ππ\pi\pi scattering. Unlike the large-Nc{N_{c}} limit, there is no reason to expect an enhancement of the axial couplings, which in that case gives rise to the contracted spin-flavor symmetries Dashen:1993as .

4 Discussion

In QCD the number of colors, Nc{N_{c}}, is a parameter that appears in the action and in some sense acts as a knob that dials the amount of quantum correlation in the hadronic SS-matrix. The simplifications, counting rules and enhanced symmetries implied by the large-Nc{N_{c}} approximation have proved highly successful in explaining regularity in the hadronic spectrum. Recent work in Ref. Beane:2018oxh has conjectured that, independent of the value of Nc{N_{c}}, quantum entanglement is minimized in hadronic SS-matrices. Verifying this conjecture relies on finding consequences of the conjecture that are distinct from large-Nc{N_{c}} predictions, and indeed this has been found to be the case in baryon-baryon scattering. In particular, minimization of entanglement near threshold leads to enhanced symmetry that is verified by lattice QCD simulations. Here this conjecture has been considered for ππ\pi\pi and πN\pi N scattering. As shown long ago by Weinberg, the scattering of soft pions off any target is completely determined by chiral symmetry PhysRevLett.17.616 and is weak at low energies. Here it has been found that the only ππ\pi\pi or πN\pi N SS-matrix, consistent with the low energy theorems, that does not entangle isospin is the identity i.e. no scattering. In the context of chiral perturbation theory this corresponds to FπF_{\pi} being large when entanglement is minimized, consistent with large Nc{N_{c}} scaling. Unlike in the large Nc{N_{c}} limit, entanglement minimization of the SS-matrix says nothing about the scaling of the baryon masses and axial couplings and therefore implies no new symmetries in the πN\pi N sector. Because of the weakness of pion processes implied by chiral symmetry, it may be the case that only systems without external Goldstone bosons (like NNNN) give non-trivial constraints from entanglement minimization. Considering general meson-nucleon scattering, it is clear that scalar-isoscalar mesons have no spin or isospin to entangle. Insofar as resonance saturation is effective, entanglement minimization would then predict the contribution to baryon-baryon scattering from the exchange of non scalar-isoscalar resonances to sum together to give an equal contribution to all spin-isospin channels Epelbaum_2002 . This would then naturally lead to the SU(16)SU(16) symmetry seen in the three flavor baryon sector Beane:2018oxh .

Techniques which make use of entanglement minimization to select out physically-relevant states and operators from an exponentially large space have a long history. For instance, tensor methods and DMRG crucially rely on the fact that ground states of reasonable Hamiltonians often exhibit much less entanglement than a typical state Eisert_2010 . In nuclear physics it has recently been shown that entanglement is a useful guiding principle when constructing many-body wave-functions of atomic nuclei Robin:2020aeh . The use of entanglement minimization to constrain hadronic SS-matrices in other contexts has also been investigated recently. The authors of Bose:2020cod ; 10.21468/SciPostPhys.9.5.081 have considered entanglement minimization as an ingredient in an effort to revive the SS-matrix bootstrap program. When applied to the ππ\pi\pi SS-matrix a correspondence is found between minima of entanglement and linear Regge trajectories. This is an intriguing prospect and may be related to the observation made here that, at least in p-wave ππ\pi\pi scattering, the entanglement power has a zero at resonance. Outside of hadronic physics, it was shown recently that the minimization of spin entanglement in scattering due to the exchange of gravitons picks out parameters which correspond to minimally coupled gravity Aoude_2020 .

An interesting and directly related line of inquiry is the connection between entanglement and renormalization group (RG) flow. As a zeroth order observation, macroscopic objects are distinguishable from their surroundings despite being coupled to the environment. Therefore, in some sense classical objects behave like coherent quantum states whose entropy does not increase when they interact with an open quantum system PhysRevLett.70.1187 . This “motivates” the idea that at large scales a fixed point of the entanglement entropy is reached. From an RG point of view this may be manifest in the entanglement structure between different momentum modes of fields. Work has been done on computing the momentum space entanglement in both scattering events and between regions of the ground state of a quantum field theory Peschanski_2016 ; Seki_2015 ; Balasubramanian_2012 . It is speculated that the RG flow of parameters in an effective action is driven by entanglement between the IR and UV. Along a similar vein, recent work has employed numerical methods to study the ground state entanglement structure between disjoint regions of massless free scalar field theory Klco_2021 ; klco2021entanglement . In tension with the tenets of EFT, it was found that long distance entanglement gets most of its support from short distance field modes. With this in mind it may be possible that EFT, with its insensitivity to physics at the cutoff, is not the best framework with which to study entanglement. There is clearly much to explore on the relationship between entanglement, RG flow and EFT.

Acknowledgments

We would like to thank Natalie Klco and Martin J. Savage for valuable discussions. This work was supported by the U. S. Department of Energy grants DE-FG02-97ER-41014 (UW Nuclear Theory) and DE-SC0020970 (InQubator for Quantum Simulation).

References